Next Article in Journal
Pulse Design of Constant Strain Rate Loading in SHPB Based on Pulse Shaping Technique
Previous Article in Journal
Modulation of the Structural, Magnetic, and Dielectric Properties of YMnO3 by Cu Doping
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Round-the-Clock Adsorption–Degradation of Tetracycline Hydrochloride by Ag/Ni-TiO2

1
College of Chemistry & Environment, Southwest Minzu University, Chengdu 610225, China
2
Department of Chemistry, College of Science, King Saud University, Riyadh 11451, Saudi Arabia
3
School of Physical Science and Technology, Soochow University, Suzhou 215006, China
*
Author to whom correspondence should be addressed.
Materials 2024, 17(12), 2930; https://doi.org/10.3390/ma17122930
Submission received: 11 May 2024 / Revised: 6 June 2024 / Accepted: 12 June 2024 / Published: 14 June 2024
(This article belongs to the Special Issue Advanced Materials for Battery Applications and Photoelectric Devices)

Abstract

:
The synergy of adsorption and photocatalysis is a good method to remove organic pollutants in wastewater. In recent decades, persistent photocatalysis has gained considerable interest for its ability to sustain the catalytic degradation of organic pollutants in the dark. Herein, we report three different TiO2 nanomaterials to remove tetracycline hydrochloride (TCH) in solution. We found that the removal ability of TiO2, Ni-TiO2, and Ag/Ni-TiO2 is 8.8 mg/g, 13.9 mg/g and 23.4 mg/g, respectively, when the initial concentration of TCH is 50 mg/L. Chemical adsorption could be the rate-determining step in the TCH adsorption process. Moreover, Ag nanoparticles dispersed on Ni doped TiO2 surface act as traps to capture photo-generated electrons upon illumination with indoor light. The holes in Ag/Ni-TiO2 serve as critical oxidative species in TCH degradation under dark conditions. This work provides new insights into the design of persistent photocatalysts that can be activated by weak illumination and degrade organic pollutants in wastewater after sunset.

1. Introduction

Tetracycline hydrochloride (TCH) is a typical antibiotic and has long been used as a broad-spectrum antibacterial agent in the livestock, poultry, and human medical industries [1,2]. It is obtained from a common hydronaphthacene core comprising four fused rings and is difficult to degrade due to its stable chemical structure [3,4]. TCH excreted into the aqueous environment through human and animal urine, causing harm to the ecological environment and human health, has raised significant concerns [5,6]. Therefore, an efficient method is urgently needed to remove TCH residues from the environment.
Adsorption is considered as an appropriate and feasible method to remove TCH from the aqueous environment due to its economic feasibility and convenience [2,7]. Zhang et al. prepared a biochar–ceramsite composite and obtained a 30.72 mg/g adsorption capacity of TCH [7]. Kong et al. found that an MOF-525(Co) metal organic framework showed a 469.5 mg/g adsorption capacity of TCH at 298 K [6]. However, TCH enriched with adsorption materials must still be degraded. Photocatalysis is a highly efficacious way to solve this problem because of its high efficiency, deep oxidation capability, good sustainability, cost effectiveness, etc. [8]. Photogenerated reactive oxidative species (h+, ·O2, ·OH, H2O2, etc.) could degrade organic pollutants and avoid secondary pollution [9]. Recently, the combination of adsorption and photocatalysis has been extensively investigated for the removal and degradation of organic pollutants in water, leveraging their complementary advantages [10,11,12]. S-defected MoS2 coupling with S,N-codoped porous biochar prepared by the Zheng group showed great adsorption capacity and a 97.99% removal efficiency of TCH within 120 min [12]. Lu et al. prepared a TiO2/(Bi2O3/Bi2O2.33) heterostructure which showed an excellent adsorption and removal efficiency of TCH [13]. Although the above-reported materials can remove TCH by synergistic adsorption–photocatalysis, a continuous supply of light is essential for the degradation of TCH. Once the irradiation stops, the photocatalyst will immediately become inactive [14].
Persistent photocatalysis is urgently needed to be developed for the application of TCH degradation during day and night. Persistent photocatalysis is mainly constructed by heterojunction systems, which are composed of a charge storage material and a photosensitive material for charge injection [14,15,16,17]. The charge storage material can store the excess photogenerated energy (h+ or e) from the photosensitive material with light excitation and then release the stored energy without light [9]. Wang et al. prepared a multifunctional catalyst of T/NC/MoS2@Ag NFs which removed 97.2% tetracycline under weak irradiation by released electrons [18]. Fu et al. found that the electron storage capacity of Ti3C2/TiO2/Ag was up to 0.125 μmol/g, which resulted in an improved dark-reaction activity to degrade organic pollutants [15].
TiO2 is vigorously applied as a photosensitive material to photocatalysis owning to its excellent stability and low cost. Nonetheless, the development of TiO2 is hindered by its wide band gap of 3.2 eV and poor utilization of visible light [19]. Doping with metals (Ni, Cr, Fe, Co, etc.) into the TiO2 lattice would actively modify the TiO2 physical properties because of the creation of an impurity energy level [20,21,22]. Au, Pt, and Ag nanoparticles have capacitive properties that can store electrons produced in the light-reaction stage [4,17,22,23]. Moreover, a Schottky junction, formed at the metal/semiconductor interface, is a typical heterojunction [24]. It can induce the formation of surface sub-band-gap states, significantly affect the charge separation process, redistribute interfacial charges, and modulate the chemisorption of reaction intermediates [25,26].
Herein, we synthesize a new Ag/Ni-TiO2 composite for the efficient persistent photocatalysis of TCH in the absence of in situ irradiation light sources. Ni-doped TiO2, which could be activated by visible light, was selected as the photosensitive material. Ag nanoparticles were used as the trap center to store photogenerated charge carriers upon illumination. We found that chemical adsorption limited the removal efficiency of TCH during the adsorption process. Ag/Ni-TiO2 could be activated by weak indoor light to generate photoinduced holes as critical oxidative species in TCH degradation under dark.

2. Materials and Methods

2.1. Preparation of Materials

TiO2 was prepared using a hydrothermal method according to the literature [27,28]. Initially, 0.01 mol titanium sulfate (Sinopharm Chemical Reagent Co., Shanghai, China) was added to 60 mL of deionized water and stirred for 20 min. Subsequently, the mixture was transferred into a 100 mL Teflon-lined stainless-steel autoclave and sealed. After heating at 180 °C for 12 h, the precursor was separated by filtration and washed with H2O, and then dried at 60 °C in a vacuum oven for 12 h.
For the synthesis of Ni-TiO2, 0.01 mol titanium sulfate and 0.0025 mol nickel nitrate hexahydrate (Chron Chemicals, Chengdu, China) were simultaneously added to 60 mL of deionized water, and the subsequent treatment process was the same as that for the preparation of TiO2 mentioned above.
For the synthesis of Ag/Ni-TiO2, 100 mg Ni-TiO2 and 625 μL AgNO3 solution (0.1 mol/L) (Aladdion Biochemical Technology Co., Shanghai, China) were simultaneously added to 20 mL of deionized water, followed by ultrasonication for 30 min to form mixture A. Meanwhile, 0.047 g NaBH4 (Chron Chemicals, Chengdu, China) and 0.066 g Na2CO3 (Sinopharm Chemical Reagent Co., Shanghai, China) were simultaneously added to 6 mL of deionized water and stirred for 20 min to form mixture B. Then, mixture B was slowly added into mixture A in an ice-water bath under stirring. Finally, the precursor was separated by filtration and washed with H2O, and then dried at 60 °C in a vacuum oven for 12 h. The synthetic route is shown in Figure 1.

2.2. Characterization

X-ray diffraction (XRD) patterns were recorded on a Smartlab X-ray diffractometer (Rigaku, Tokyo, Japan) with Cu Kα radiation (λ = 0.15405 nm). The determination of specific surface area and pore volumes involved the utilization of N2 adsorption and desorption isotherms, which were conducted using a Micromeritics ASAP 2460 instrument (Micromeritics, Atlanta, GA, USA). The morphology of the samples was characterized using scanning electron microscopy (SEM, Hitachi S-4800) (Hitachi, Tokyo, Japan). X-ray photoelectron spectroscopy (XPS) was performed on a Thermo-Scientific K-Alpha (Thermo Fisher Scientific, Waltham, MA, USA). The content of Ni in Ni-TiO2 was determined using inductively coupled plasma-atomic emission spectrometry (ICP-AES, Agilent 720ES) (Agilent, Santa Clara, CA, USA). Fourier-transform infrared spectroscopy (FT-IR) was performed using an Agilent Cary 660 (Agilent, Santa Clara, CA, USA).

2.3. Adsorption and Degradation Experiment

2.3.1. Adsorption–Degradation Performance of TCH

A total of 50 mg of each sample was added to a 50 mL TCH solution with different concentrations (10 mg/L, 30 mg/L, and 50 mg/L), and the adsorption experiments were performed in a crystal vessel (100 mL) at 293 K with continuous stirring. Samples were gathered at regular intervals and the concentration of TCH was assessed. This experiment was carried out in the dark.

2.3.2. Adsorption–Degradation Cycles of TCH over Ag/Ni-TiO2

A total of 50 mg of Ag/Ni-TiO2 was added to a 10 mg/L aqueous TCH solution (50 mL), followed by stirring under dark for 150 min at 293 K. Then, the Ag/Ni-TiO2 powders were collected by high-speed centrifugalizing and vacuum drying at 60 °C, and then added to the TCH solution again.

2.3.3. Adsorption–Degradation Cycles of TCH over Illuminated Materials

A total of 50 mg of Ag/Ni-TiO2 (or Ni-TiO2) was added to a 10 mg/L aqueous TCH solution (50 mL), followed by stirring in the dark for 150 min at 293 K. Then, the Ag/Ni-TiO2 (or Ni-TiO2) powders were collected by high-speed centrifugalizing and vacuum drying at 60 °C. They were then illuminated using a 350 W Xe lamp (or Xe lamp with a wavelength exceeding 420 nm) for 2 h and then added to the TCH solution again.

2.4. Determination of TCH

The concentration of TCH was measured using a UV-Vis spectrometer at 358 nm [6]. The removal efficiency and adsorption capacity (qt) for TCH were calculated according to Equations (1) and (2):
Removal   efficiency = ( C 0 C t ) C 0 × 100 %
q t = V ( C 0 C t ) m
where the initial concentration is denoted as C0, and the concentration at time t for TCH is denoted as Ct. V represents the initial volume of the TCH solution (50 mL), and m denotes the dosage of the samples (50 mg).

3. Results and Discussion

The crystal phases of TiO2, Ni-TiO2, and Ag/Ni-TiO2 are shown in Figure 2a. TiO2 and Ni-TiO2 were characterized by the anatase phase (JPCDS No. 21-1272) [29]. The doping of Ni did not alter the crystal phase structure of TiO2. The other five peaks centered at 38.1°, 44.3°, 64.4°, 77.4°, and 81.5° were identified as the (111), (200), (220), (311), and (222) planes corresponding to Ag (JPCDS No. 89-3722) over the Ag/Ni-TiO2 sample [30]. The results suggested that Ag nanoparticles dispersed on the surface of Ni-TiO2.
N2 adsorption and desorption isotherms were conducted to analyze the porosity and specific surface area of samples. As shown in Figure 2b, all samples exhibited type Ⅳ isotherm which signifies the mesoporous nature of the samples [19]. The specific surface area of Ni-TiO2 was 128.34 m2/g, exceeding that of pure TiO2 at 118.77 m2/g. And the specific surface area of Ag/Ni-TiO2 was 113.90 m2/g, which indicates that pores on Ni-TiO2 were partially filled or blocked by Ag nanoparticles. The average pore diameters of TiO2, Ni-TiO2, and Ag/Ni-TiO2 were 7.06, 5.74, and 8.42 nm (Figure 2c), respectively, suggesting that all samples exhibited a mesoporous structure.
Figure 3a–d show the SEM images of TiO2, Ni-TiO2, and Ag/Ni-TiO2, respectively. It is evident that the agglomeration of nanoparticles occurred on the TiO2 sample, as shown in Figure 3a. The doping of Ni promoted the dispersion of nanoparticles in Figure 3b. The loading of Ag nanoparticles did not significantly affect the microstructure in Figure 3c,d. To verify the composition of Ag/Ni-TiO2, we carried out energy dispersive spectroscopic (EDS) mapping, as shown in Figure 3e–h. Chemical mapping affirms the presence of Ag, Ti, and O. According to the ICP-MS results, the content of Ni is 0.075 wt.% in Ni-TiO2, which results in a weak Ni mapping signal.
The XPS spectra of all samples are shown in Figure 4. The wide survey XPS spectra of TiO2, Ni-TiO2, and Ag/Ni-TiO2 in Figure S1 indicate the presence of C, O, Ti, and Ag elements. O 1s spectra (Figure 4a) were deconvoluted into two peaks at 531.5–531.1 and 530.0–529.7 eV, assigned to adsorbed oxygen and lattice oxygen, respectively [31]. For Ti 2p, the splitting peaks at 464.4–464.2 and 458.8–454.4 eV, respectively, correspond to Ti 2p1/2 and Ti 2p3/2 of Ti4+ [11]. The Ni 2p signal is almost not detected over Ni-TiO2 and Ag/Ni-TiO2 in Figure 4c. However, for Ni-TiO2, the Ti 2p3/2 and lattice oxygen peaks displayed a negative shift of 0.2 eV and 0.1 eV, respectively, compared with pure TiO2, which was attributed to the perturbation of the doped Ni atoms in the crystal structure of TiO2 [32].
The Ag 3d XPS spectrum of Ag/Ni-TiO2 exhibits distinct peaks at 373.4 eV and 367.4 eV, corresponding to the Ag 3d3/2 and Ag 3d5/2 states, respectively. The binding gap between the peaks is 6 eV, which aligns with the Ag0 state [33]. Additionally, it should be mentioned that there is a negative shift of approximately −0.8 eV in the binding energy from 374.2 to 373.4 eV in comparison with pristine Ag. This could be attributed to the intimate contact established between Ag nanoparticles and Ni-TiO2, resulting in the formation of a Schottky junction, with some electrons being transferred across the barrier from Ni-TiO2 and trapped in Ag particles [34,35,36]. For Ag/Ni-TiO2, the Ti 2p3/2 and lattice oxygen peaks displayed a negative shift of 0.2 eV and 0.2 eV, respectively, compared with Ni-TiO2. Such shifts in the binding energy could be due to the adjustment of the Fermi level in the composite Ag/Ni-TiO2 [31,37].
Figure 5 displays the adsorption–degradation experiments of TCH over the samples under dark conditions. The removal efficiency of TCH by TiO2, Ni-TiO2, and Ag/Ni-TiO2 was 42.77%, 55.11%, and 81.03% at 30 min, and then reached 58.93%, 65.59%, and 87.89% at 150 min when the initial concentration of TCH was 10 mg/L, as shown in Figure 5a. It is evident that the removal efficiency decreased when the initial concentration of TCH increased to 30 and 50 mg/L over all samples, as shown in Figure 5b,c, and Ag/Ni-TiO2 performed optimum adsorption–degradation.
Further, the adsorption capacity (qt) of TCH over the three samples was calculated and is displayed in Figure 5d–f. The fast adsorption occurring within the initial 30 min could be attributed to the ample adsorption sites on the surface of the samples [6]. As shown in Figure 5d, qt reached 5.9, 8.1, and 8.8 mg/g at 150 min when the concentration of TCH was 10, 30, and 50 mg/L, respectively, which indicates that the adsorption equilibrium was reached when TCH concentration was 30 mg/L on TiO2. However, qt kept increasing over Ni-TiO2 and Ag/Ni-TiO2 when the concentration of TCH increased, and reached 13.9 and 23.4 mg/g, respectively, at 150 min when the concertation of TCH was 50 mg/L (Figure 5e,f). The above results indicate that the doping of Ni atoms and Ag nanoparticles strengthens the removal performance of TiO2.
To study the adsorption kinetics during TCH removal, the removal efficiency over the Ag/Ni-TiO2 composite between 0 and 40 min was tested. The adsorption kinetics were well matched with the pseudo-second-order kinetic model (PSOK) due to higher correlation coefficient (R2) values, as shown in Figure S2. Therefore, the PSOK model is used to describe the adsorption kinetics over three samples. As shown in Figure 5d–f, the R2 values for all samples are approaching 1.0, which suggests that the chemical adsorption could be the rate-determining step in the TCH adsorption process over all samples [38,39,40]. Moreover, according to the BET results in Figure 2b, Ag/Ni-TiO2 has the smallest specific surface area, which further demonstrates that physical adsorption does not limit the removal of TCH.
As shown in Figure 6a, the composition of fresh and TCH-adsorbed samples was analyzed using FT-IR tests. The peak at 695 cm−1 corresponds to the stretching vibration of Ti-O-Ti [40]. The adsorption peaks at 1626 and 3200 cm−1 are attributed to -OH and H2O, respectively [22]. It can be seen that adsorbed CO2 at 2358 cm−1 was detected over TCH-adsorbed samples [41], which indicates that the adsorbed TCH can undergo dark degradation, resulting in the production of CO2.
In order to further clarify the above proposal, free radical capture experiments in the dark were carried out. Isopropyl alcohol (IPA), 1,4-Benzoquinone (PBQ), and ethylenediaminetetraacetic acid sodium (EDTA-2Na) were used as scavengers for hydroxyl radicals, superoxide radicals, and holes, respectively [11,42]. As shown in Figure 6b, over the TiO2 sample, after adding IPA, PBQ, and EDTA-2Na, the removal efficiency of TCH decreased from 58% to 41%, 37%, and 55%, respectively. The results indicate that hydroxyl radicals and superoxide radicals, which could be formed from the active oxygen in the solution [43], are the main active species to degrade TCH to CO2 over the TiO2 surface. Over the Ni-TiO2 sample, after adding IPA, PBQ, and EDTA-2Na, the removal efficiency of TCH decreased from 65% to 40%, 40%, and 55%, respectively, which suggests that the doping of Ni atoms could promote the generation of holes to degrade TCH. Over the Ag/Ni-TiO2 sample, after adding IPA, PBQ, and EDTA-2Na, the removal efficiency of TCH decreased from 88% to 78%, 86%, and 53%, respectively, which suggests that holes generated on the surface of Ag/Ni-TiO2 are the main active species leading to the removal of TCH in the dark. It is worth noting that all samples were only irradiated by an indoor light during the preparation process.
To better demonstrate the removal ability of Ag/Ni-TiO2, an adsorption–degradation cycle of TCH was conducted for five times, as exhibited in Figure 6c. The removal efficiency of TCH reached 84%, 56%, 42%, 25%, and 25% in the first, second, third, fourth, and fifth process, respectively. The results suggest that the adsorption capacity reached saturation after the third iteration. The removal efficiency of 25% in the fourth and fifth processes could be caused by the degradation of TCH by holes over Ag/Ni-TiO2. The XRD pattern, SEM image, and EDS mapping of cycled-Ag/Ni-TiO2 are shown in Figures S3–S5. It can be seen that the crystal phase and micromorphology of the composite remained unchanged. The results indicate that the Ag/Ni-TiO2 composite had excellent stability during the adsorption–degradation cycle of TCH.
Further, the cycled Ag/Ni-TiO2 was irradiated with a Xe lamp for 2 h and then placed into the TCH solution again for the degradation application in the dark. As shown in Figure 6d, the removal efficiency of TCH remained at about 95% during five cycle processes in the dark. When the cycled Ag/Ni-TiO2 was irradiated using a Xe lamp with a wavelength exceeding 420 nm for 2 h, as shown in Figure S6, the removal efficiency of TCH degraded to 85% in the fourth cycle, which could be attributed to the loss of Ag nanoparticles on the surface. Furthermore, an adsorption–degradation cycle of TCH over Ni-TiO2 was conducted, as shown in Figures S7 and S8. The removal efficiency of TCH increased to 90% during the cycles when the cycled Ni-TiO2 powder was irradiated with a Xe lamp for 2 h. When the cycled Ni-TiO2 powder was irradiated using a Xe lamp with a wavelength exceeding 420 nm for 2 h, the removal efficiency of TCH remained at about 70% during the cycles. The above results indicated that Ni-TiO2 and Ag/Ni-TiO2 could be activated by pre-illumination with visible light to store enough charge for the effective and persistent degradation of TCH under dark conditions. UV light illumination could enhance the stored charge on the surface of the composite. The higher performance of Ag/Ni-TiO2 could be attributed to the formation of a Schottky junction at the interface according to the XPS results. This junction facilitates the transfer of excited electrons to the Ag nanoparticles because of the silver’s higher work function, effectively boosting the efficiency of charge separation and leaving energetic positive charges in Ni-TiO2, which essentially function as holes for the oxidation of TCH in the dark [22,44,45].

4. Conclusions

In summary, we have reported the excellent ability of a Ag/Ni-TiO2 composite for the persistent photodegradation of TCH with good recycling capability. The removal capacity of TCH reached 23.4 mg/g in the presence of an indoor light-irradiated Ag/Ni-TiO2 composite. We found that adsorption and degradation simultaneously occurred during TCH removal and chemical adsorption could be the rate-determining step in the TCH adsorption process. The loaded Ag nanoparticles act as trapping sites to capture photo-generated electrons upon illumination. The holes in Ag/Ni-TiO2 served as critical oxidative species in TCH degradation in the dark. This work provides new insights into the design of persistent photocatalysts that can be activated by weak visible light and degrade organic pollutants after sunset.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma17122930/s1, Figure S1: Survey XPS spectra of TiO2, Ni-TiO2, and Ag/Ni-TiO2; Figure S2: Nonlinear regressions of the pseudo-first-order kinetic model (PFOK) (a) and pseudo-second-order kinetic model (PSOK) (b) over Ag/Ni-TiO2 between 0 and 40 min; Figure S3: XRD of Ag/Ni-TiO2 and cycled-Ag/Ni-TiO2; Figure S4: SEM of cycled-Ag/Ni-TiO2; Figure S5: (a) SEM image of cycled-Ag/Ni-TiO2 at low magnification, (b) EDS Ag mapping of the region shown in (a), (c) EDS Ni mapping of the region shown in (a), (d) EDS Ti mapping of the region shown in (a), (e) EDS O mapping of the region shown in (a); Figure S6: Adsorption–degradation cycles of TCH under dark condition, the cycled-Ag/Ni-TiO2 was irradiated using a Xe lamp with a wavelength exceeding 420 nm for 2 h; Figure S7: Adsorption–degradation cycles of TCH under dark condition, the cycled-Ni-TiO2 was irradiated using a Xe lamp for 2 h; Figure S8: Adsorption–degradation cycles of TCH under dark condition, the cycled-Ni-TiO2 was irradiated using a Xe lamp with a wavelength exceeding 420 nm for 2 h.

Author Contributions

S.M.: Methodology, Data curation, and Investigation. Y.Q.: Methodology and Data curation. K.S.: Methodology. J.A.: Supervision. W.T.: Supervision. Z.M.: Conceptualization, Supervision, Project administration, Writing—Original Draft, and Funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Fundamental Research Funds for the Central Universities, Southwest Minzu University (Grant No. RQD2021085), and the Natural Science Foundation of Sichuan Province (Grant No. 2024NSFSC1008). The authors are also thankful to Researcher Supporting Project (number RSP2024R391), King Saud University, Riyadh, Saudi Arabia.

Data Availability Statement

The original contributions presented in the study are included in the article and Supplementary Materials, further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Nie, Y.; Zhao, C.; Zhou, Z.; Kong, Y.; Ma, J. Hydrochloric acid-modified fungi-microalgae biochar for adsorption of tetracycline hydrochloride: Performance and mechanism. Bioresour. Technol. 2023, 383, 129224. [Google Scholar] [CrossRef] [PubMed]
  2. Zhang, W.; Bian, Z.; Peng, Y.; Tang, H.; Wang, H. Dual-function oxygen vacancy of BiOBr intensifies pollutant adsorption and molecular oxygen activation to remove tetracycline hydrochloride. Chem. Eng. J. 2023, 451, 138731. [Google Scholar] [CrossRef]
  3. Xu, L.; Zhang, H.; Xiong, P.; Zhu, Q.; Liao, C.; Jiang, G. Occurrence, fate, and risk assessment of typical tetracycline antibiotics in the aquatic environment: A review. Sci. Total Environ. 2021, 753, 141975. [Google Scholar] [CrossRef] [PubMed]
  4. Liu, Y.; Wang, L.; Dai, X.; Zhang, J.; Li, J.; Ma, Y.; Han, Q.; Dong, Y. Research on the adsorption-photocatalytic synergistic degradation of tetracycline by Au nanoparticles/TiO2 nanorods/biochar. J. Alloys Compd. 2024, 976, 172985. [Google Scholar] [CrossRef]
  5. Zhang, T.; Liu, Y.; Rao, Y.; Li, X.; Yuan, D.; Tang, S.; Zhao, Q. Enhanced photocatalytic activity of TiO2 with acetylene black and persulfate for degradation of tetracycline hydrochloride under visible light. Chem. Eng. J. 2020, 384, 123350. [Google Scholar] [CrossRef]
  6. Kong, Y.; Lu, H.; Wang, R.; Yang, Q.; Huang, B.; Zhou, Q.; Hu, W.; Zou, J.; Chen, Q. Adsorption characteristics of tetracycline hydrochloride and oxytetracycline by a MOF-525(Co) metal organic framework. Colloids Surf. A Physicochem. Eng. Asp. 2023, 677, 132443. [Google Scholar] [CrossRef]
  7. Zhang, K.; Zhang, L.; Dong, X.; Zhao, Y.; Li, F.; Cen, Q. Efficient adsorption of tetracycline hydrochloride on biochar-ceramsite composite: Optimization of response surface methodology and investigation of adsorption mechanism. Mater. Today Sustain. 2023, 24, 100525. [Google Scholar] [CrossRef]
  8. Wang, C.; Yan, R.; Cai, M.; Liu, Y.; Li, S. A novel organic/inorganic S-scheme heterostructure of TCPP/Bi12O17Cl2 for boosting photodegradation of tetracycline hydrochloride: Kinetic, degradation mechanism, and toxic assessment. Appl. Surf. Sci. 2023, 610, 155346. [Google Scholar] [CrossRef]
  9. Zhang, C.; Xiong, W.; Li, Y.; Lin, L.; Zhou, X.; Xiong, X. Continuous inactivation of human adenoviruses in water by a novel g-C3N4/WO3/biochar memory photocatalyst under light-dark cycles. J. Hazard. Mater. 2023, 442, 130013. [Google Scholar] [CrossRef] [PubMed]
  10. Xue, B.; Yang, C.; Chang, M.; Liu, D. Structural design of core-shell ZIF-8@NH2-MIL-125 photocatalyst for synergistic adsorption-photocatalytic degradation of tetracycline hydrochloride. J. Alloys Compd. 2024, 973, 172850. [Google Scholar] [CrossRef]
  11. Cheng, X.; Guan, R.; Chen, Y.; Qian, Y.; Shang, Q.; Sun, Y. Adsorption and photocatalytic degradation process of oxytetracycline using mesoporous Fe-TiO2 based on high-resolution mass spectrometry. Chem. Eng. J. 2023, 460, 141618. [Google Scholar] [CrossRef]
  12. Peng, H.; Wang, L.; Zheng, X. Efficient adsorption-photodegradation activity of MoS2 coupling with S,N-codoped porous biochar derived from chitosan. J. Water Process Eng. 2023, 51, 103426. [Google Scholar] [CrossRef]
  13. Lu, R.; Zhao, S.; Yang, Y.; Wang, Y.; Huang, H.; Hu, Y.; Rodriguez, R.; Chen, J. Efficient adsorption and photocatalytic degradation of organic contaminants using TiO2/(Bi2O3/Bi2O2.33) nanotubes. Mater. Today Chem. 2023, 32, 101641. [Google Scholar] [CrossRef]
  14. Zhang, C.; Li, Y.; Li, M.; Shuai, D.; Zhou, X.; Xiong, X.; Wang, C.; Hu, Q. Continuous photocatalysis via photo-charging and dark-discharging for sustainable environmental remediation: Performance, mechanism, and influencing factors. J. Hazard. Mater. 2021, 420, 126607. [Google Scholar] [CrossRef] [PubMed]
  15. Fu, X.; Kong, Y.; Wang, M.; Cai, T.; Zeng, Q. MXene derived Ti3C2/TiO2/Ag persistent photocatalyst with enhanced electron storage capacity for round-the-clock degradation of organic pollutant. J. Colloid Interface Sci. 2024, 656, 233. [Google Scholar] [CrossRef] [PubMed]
  16. Zhou, D.-M.; Chen, L.-J.; Zhao, X.; Yan, L.-X.; Yan, X.-P. Persistent production of multiple active species with copper doped zinc gallate nanoparticles for light-independent photocatalytic degradation of organic pollutants. J. Colloid Interface Sci. 2024, 668, 540. [Google Scholar] [CrossRef] [PubMed]
  17. Loh, J.Y.Y.; Sharma, G.; Kherani, N.P.; Ozin, G.A. Post-illumination photoconductivity enables extension of photo-catalysis after sunset. Adv. Energy Mater. 2021, 11, 2101566. [Google Scholar] [CrossRef]
  18. Wang, C.; Li, Y.; Fan, M.; Yu, X.; Ding, J.; Yan, L.; Qin, G.; Yang, J.; Zhang, Y. Controllable growth of silver nanoparticles on titanium dioxide/nitrogen-doped carbon nanofiber/molybdenum disulfide: Toward enhanced photocatalytic-activating peroxymonosulfate performance and “memory catalysis”. Chem. Eng. J. 2024, 479, 147752. [Google Scholar] [CrossRef]
  19. Liang, C.; Li, C.; Zhu, Y.; Du, X.; Zeng, Y.; Zhou, Y.; Zhao, J.; Li, S.; Liu, X.; Yu, Q.; et al. Light-driven photothermal catalysis for degradation of toluene on CuO/TiO2 Composite: Dominating photocatalysis and auxiliary thermalcatalysis. Appl. Surf. Sci. 2022, 601, 154144. [Google Scholar] [CrossRef]
  20. Shaban, M.; Ahmed, A.M.; Shehata, N.; Betiha, M.A.; Rabie, A.M. Ni-doped and Ni/Cr co-doped TiO2 nanotubes for enhancement of photocatalytic degradation of methylene blue. J. Colloid Interface Sci. 2019, 555, 31. [Google Scholar] [CrossRef]
  21. Kaur, N.; Verma, A.; Thakur, I.; Basu, S. In-situ dual effect of Ag-Fe-TiO2 composite for the photocatalytic degradation of Ciprofloxacin in aqueous solution. Chemosphere 2021, 276, 130180. [Google Scholar] [CrossRef] [PubMed]
  22. Fu, C.; Liu, L.; Zhao, G. Role of Ag Loading on the concentration of surface-reaching photoexcited holes in TiO2 nanoparticles. J. Phys. Chem. C 2024, 128, 810. [Google Scholar] [CrossRef]
  23. Septiningrum, F.; Yuwono, A.H.; Maulana, F.A.; Nurhidayah, E.; Dhaneswara, D.; Sofyan, N.; Hermansyah, H.; Purwanto, W.W. Mangosteen pericarp extract mediated synthesis of Ag/TiO2 nanocomposite and its application on organic pollutant degradation by adsorption-photocatalytic activity. Curr. Res. Green Sustain. Chem. 2024, 8, 100394. [Google Scholar] [CrossRef]
  24. Li, X.; Shen, R.; Ma, S.; Chen, X.; Xie, J. Graphene-based heterojunction photocatalysts. Appl. Surf. Sci. 2018, 430, 53. [Google Scholar] [CrossRef]
  25. Liu, Y.; Guo, J.; Zhu, E.; Liao, L.; Lee, S.-J.; Ding, M.; Shakir, I.; Gambin, V.; Huang, Y.; Duan, X. Approaching the Schottky–Mott limit in van der Waals metal–semiconductor junctions. Nature 2018, 557, 696. [Google Scholar] [CrossRef]
  26. Monteblanco, E.; Donatini, F.; Hehn, M.; Lacour, D.; Lassailly, Y.; Peretti, J.; Rougemaille, N. Transmission of high-energy electrons through metal-semiconductor Schottky junctions. Phys. Rev. B 2019, 100, 205301. [Google Scholar] [CrossRef]
  27. Liu, S.; Liu, C.; Wang, W.; Cheng, B.; Yu, J. Unique photocatalytic oxidation reactivity and selectivity of TiO2–graphene nanocomposites. Nanoscale 2012, 4, 3193. [Google Scholar] [CrossRef] [PubMed]
  28. Deng, H.; Cheng, Y.; Li, J.; Zhang, Q.; Cao, Y.; Wang, X. Synergistic effect of surfactant hexadecyltrimethylammonium supporting and hydrous titania impregnation onto zeolite in enhanced removal of anionic X-3B dye from aqueous solution. Environ. Prog. Sustain. Energy 2022, 41, e13843. [Google Scholar] [CrossRef]
  29. Paul, T.C.; Podder, J. Synthesis and characterization of Zn-incorporated TiO2 thin films: Impact of crystallite size on X-ray line broadening and bandgap tuning. Appl. Phys. A 2019, 125, 818. [Google Scholar] [CrossRef]
  30. Liu, Y.; Zhang, Q.; Xu, M.; Yuan, H.; Chen, Y.; Zhang, J.; Luo, K.; Zhang, J.; You, B. Novel and efficient synthesis of Ag-ZnO nanoparticles for the sunlight-induced photocatalytic degradation. Appl. Surf. Sci. 2019, 476, 632. [Google Scholar] [CrossRef]
  31. Sun, Q.M.; Xu, J.J.; Tao, F.F.; Ye, W.; Zhou, C.; He, J.H.; Lu, J.M. Boosted inner surface charge transfer in perovskite nanodots@mesoporous titania frameworks for efficient and selective photocatalytic CO2 reduction to methane. Angew. Chem. Int. Ed. 2022, 61, e202200872. [Google Scholar] [CrossRef] [PubMed]
  32. Zhang, Y.; Li, Y.; Ni, D.; Chen, Z.; Wang, X.; Bu, Y.; Ao, J.P. Improvement of BiVO4 Photoanode Performance during Water Photo-Oxidation Using Rh-Doped SrTiO3 Perovskite as a Co-Catalyst. Adv. Funct. Mater. 2019, 29, 1902101. [Google Scholar] [CrossRef]
  33. Lei, Q.; Miao, R.; Li, X.; Liu, X.; Li, Y.; He, Z.; Xie, H.; Song, F.; Liu, X.; Liu, H. Efficient hydrogen production from formic acid over Ag@AgPd nanotriangulars at room temperature. Fuel 2024, 355, 129539. [Google Scholar] [CrossRef]
  34. Duan, Y.; Zhang, M.; Wang, L.; Wang, F.; Yang, L.; Li, X.; Wang, C. Plasmonic Ag-TiO2−x nanocomposites for the photocatalytic removal of NO under visible light with high selectivity: The role of oxygen vacancies. Appl. Catal. B Environ. 2017, 204, 67. [Google Scholar] [CrossRef]
  35. Jiang, J.; Zhao, P.; Shi, L.; Yue, X.; Qiu, Q.; Xie, T.; Wang, D.; Lin, Y.; Mu, Z. Insights into the interface effect in Pt@BiOI/ZnO ternary hybrid composite for efficient photodegradation of phenol and photogenerated charge transfer properties. J. Colloid Interface Sci. 2018, 518, 102. [Google Scholar] [CrossRef] [PubMed]
  36. Lu, L.; Wang, G.; Xiong, Z.; Hu, Z.; Liao, Y.; Wang, J.; Li, J. Enhanced photocatalytic activity under visible light by the synergistic effects of plasmonics and Ti3+-doping at the Ag/TiO2-heterojunction. Ceram. Int. 2020, 46, 10667. [Google Scholar] [CrossRef]
  37. Chen, P.; Du, T.; Jia, H.; Zhou, L.; Yue, Q.; Wang, H.; Wang, Y. A novel Bi2WO6/Si heterostructure photocatalyst with Fermi level shift in valence band realizes efficient reduction of CO2 under visible light. Appl. Surf. Sci. 2022, 585, 152665. [Google Scholar] [CrossRef]
  38. Revellame, E.D.; Fortela, D.L.; Sharp, W.; Hernandez, R.; Zappi, M.E. Adsorption kinetic modeling using pseudo-first order and pseudo-second order rate laws: A review. Clean. Eng. Technol. 2020, 1, 100032. [Google Scholar] [CrossRef]
  39. Kim, H.-G.; Choi, K.; Lee, K.; Lee, S.; Jung, K.-W.; Choi, J.-W. Controlling the structural robustness of zirconium-based metal organic frameworks for efficient adsorption on tetracycline antibiotics. Water 2021, 13, 1869. [Google Scholar] [CrossRef]
  40. Sui, G.; Li, J.; Du, L.; Zhuang, Y.; Zhang, Y.; Zou, Y.; Li, B. Preparation and characterization of g-C3N4/Ag–TiO2 ternary hollowsphere nanoheterojunction catalyst with high visible light photocatalytic performance. J. Alloys Compd. 2020, 823, 153851. [Google Scholar] [CrossRef]
  41. Coenen, K.; Gallucci, F.; Mezari, B.; Hensen, E.; van Sint Annaland, M. An in-situ IR study on the adsorption of CO2 and H2O on hydrotalcites. J. CO2 Util. 2018, 24, 228. [Google Scholar] [CrossRef]
  42. Ni, J.; Wang, W.; Liu, D.; Zhu, Q.; Jia, J.; Tian, J.; Li, Z.; Wang, X.; Xing, Z. Oxygen vacancy-mediated sandwich-structural TiO2−x /ultrathin g-C3N4/TiO2−x direct Z-scheme heterojunction visible-light-driven photocatalyst for efficient removal of high toxic tetracycline antibiotics. J. Hazard. Mater. 2021, 408, 124432. [Google Scholar] [CrossRef] [PubMed]
  43. Yu, C.; Lu, Y.; Zhang, Y.; Qian, A.; Zhang, P.; Tong, M.; Yuan, S. Significant contribution of solid organic matter for hydroxyl radical production during oxygenation. Environ. Sci. Technol. 2022, 56, 11878. [Google Scholar] [CrossRef] [PubMed]
  44. Wang, J.; Yu, J.; Fu, Q.; Yang, H.; Tong, Q.; Hao, Z.; Ouyang, G. Unprecedented nonphotomediated hole (h+) oxidation system constructed from defective carbon nanotubes and superoxides. ACS Cent. Sci. 2021, 7, 355. [Google Scholar] [CrossRef] [PubMed]
  45. Park, J.Y.; Baker, L.R.; Somorjai, G.A. Role of hot electrons and metal–oxide interfaces in surface chemistry and catalytic reactions. Chem. Rev. 2015, 115, 2781. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic diagram of the synthesis steps of Ag/Ni-TiO2.
Figure 1. Schematic diagram of the synthesis steps of Ag/Ni-TiO2.
Materials 17 02930 g001
Figure 2. XRD (a), N2 adsorption–desorption isotherms (b), and pore size distribution (c) of TiO2, Ni-TiO2, and Ag/Ni-TiO2.
Figure 2. XRD (a), N2 adsorption–desorption isotherms (b), and pore size distribution (c) of TiO2, Ni-TiO2, and Ag/Ni-TiO2.
Materials 17 02930 g002
Figure 3. SEM images of TiO2 (a), Ni-TiO2 (b), and Ag/Ni-TiO2 (c); (d) SEM image of Ag/Ni-TiO2 at low magnification. (e) EDS Ag mapping of the region shown in (d), (f) EDS Ni mapping of the region shown in (d), (g) EDS Ti mapping of the region shown in (d), (h) EDS O mapping of the region shown in (d).
Figure 3. SEM images of TiO2 (a), Ni-TiO2 (b), and Ag/Ni-TiO2 (c); (d) SEM image of Ag/Ni-TiO2 at low magnification. (e) EDS Ag mapping of the region shown in (d), (f) EDS Ni mapping of the region shown in (d), (g) EDS Ti mapping of the region shown in (d), (h) EDS O mapping of the region shown in (d).
Materials 17 02930 g003
Figure 4. XPS spectra of TiO2, Ni-TiO2, and Ag/Ni-TiO2. (a) O 1s, (b) Ti 2p, (c) Ni 2p, and (d) Ag 3d.
Figure 4. XPS spectra of TiO2, Ni-TiO2, and Ag/Ni-TiO2. (a) O 1s, (b) Ti 2p, (c) Ni 2p, and (d) Ag 3d.
Materials 17 02930 g004
Figure 5. Removal efficiency of TCH over TiO2, Ni-TiO2, and Ag/Ni-TiO2 for the initial concentration of 10 mg/L (a), 30 mg/L (b), and 50 mg/L (c) in the dark. Adsorption capacity of TCH and nonlinear regressions of the PSOK model over TiO2 (d), Ni-TiO2 (e), and Ag/Ni-TiO2 (f).
Figure 5. Removal efficiency of TCH over TiO2, Ni-TiO2, and Ag/Ni-TiO2 for the initial concentration of 10 mg/L (a), 30 mg/L (b), and 50 mg/L (c) in the dark. Adsorption capacity of TCH and nonlinear regressions of the PSOK model over TiO2 (d), Ni-TiO2 (e), and Ag/Ni-TiO2 (f).
Materials 17 02930 g005
Figure 6. (a) FT–IR spectrum results of the samples before and after TCH adsorption. (b) Free radical capture experiments. (c) Adsorption–degradation cycles of TCH over Ag/Ni-TiO2 under dark conditions. (d) Adsorption–degradation cycles of TCH under dark conditions; the cycled Ag/Ni-TiO2 was irradiated using a Xe lamp for 2 h.
Figure 6. (a) FT–IR spectrum results of the samples before and after TCH adsorption. (b) Free radical capture experiments. (c) Adsorption–degradation cycles of TCH over Ag/Ni-TiO2 under dark conditions. (d) Adsorption–degradation cycles of TCH under dark conditions; the cycled Ag/Ni-TiO2 was irradiated using a Xe lamp for 2 h.
Materials 17 02930 g006
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ma, S.; Qin, Y.; Sun, K.; Ahmed, J.; Tian, W.; Ma, Z. Round-the-Clock Adsorption–Degradation of Tetracycline Hydrochloride by Ag/Ni-TiO2. Materials 2024, 17, 2930. https://doi.org/10.3390/ma17122930

AMA Style

Ma S, Qin Y, Sun K, Ahmed J, Tian W, Ma Z. Round-the-Clock Adsorption–Degradation of Tetracycline Hydrochloride by Ag/Ni-TiO2. Materials. 2024; 17(12):2930. https://doi.org/10.3390/ma17122930

Chicago/Turabian Style

Ma, Siyu, Yiying Qin, Kongyuan Sun, Jahangeer Ahmed, Wei Tian, and Zhaoxia Ma. 2024. "Round-the-Clock Adsorption–Degradation of Tetracycline Hydrochloride by Ag/Ni-TiO2" Materials 17, no. 12: 2930. https://doi.org/10.3390/ma17122930

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop